Pathogenic Mechanisms in HBV-and HCV-associated Hepatocellular Carcinoma

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

REVIEWS

Pathogenic mechanisms in HBV-


and HCV-associated hepatocellular
carcinoma
Alla Arzumanyan1, Helena M. G. P. V. Reis2 and Mark A. Feitelson1
Abstract | Hepatocellular carcinoma (HCC) is a highly lethal cancer, with increasing
worldwide incidence, that is mainly associated with chronic hepatitis B virus (HBV) and/or
hepatitis C virus (HCV) infections. There are few effective treatments partly because the
cell- and molecular-based mechanisms that contribute to the pathogenesis of this tumour
type are poorly understood. This Review outlines pathogenic mechanisms that seem to be
common to both viruses and which suggest innovative approaches to the prevention and
treatment of HCC.

Hepatitis B virus (HBV) (FIG. 1) and hepatitis C virus In Africa and Asia, where HBV is endemic, 60% of HCC
Hepatitis
The accumulation of (HCV) (FIG. 2) are unrelated viruses that target the liver is associated with HBV, 20% is related to HCV and the
inflammatory cells in the liver. and that replicate in hepatocytes. Around 2 billion people remainder is distributed among other risk factors (for
are infected with HBV, and more than 350 million have example, alcohol and aflatoxin). Most treatments are
Fibrosis become chronic carriers1,2 who have persistent virus and inadequate and mortality is high. Patients with small,
The accumulation of
extracellular matrix material in
subvirus particles in their blood for more than 6 months3. single tumours are treated by surgical resection or liver
the liver that eventually forms Most infections occur at birth and more than 90% transplantation, but relapse is common9. Localized and
septa. become chronic. However, only 5–10% of adults who systemic radiation and chemotherapy have been used,
acquire infection in adulthood become carriers, but up but with limited success, as they do not target tumour-
Cirrhosis
to 30% develop progressive chronic liver disease (CLD), associated pathways and because growth-stimulatory
Occurs when fibrotic septa
completely surround islands of which appears as hepatitis, fibrosis, cirrhosis and finally pathways may override the effects of these cytotoxic
hepatocytes. hepatocellular carcinoma (HCC)4 (FIG. 3). Disease pro‑ therapies. The multi-kinase inhibitor sorafenib has
gression can arrest at any stage. The risk of developing extended the survival of patients with HCC by about
HCC among carriers with CLD ranges from tenfold to 3 months10, but there is a long way to go before thera‑
100‑fold greater compared with uninfected people2,5,6 pies that target the pathways that drive HCC are clini‑
depending on the markers and populations that are eval‑ cally available. Thus, it is timely to review the pathogenic
uated. This is one of the closest relationships between an mechanisms of HBV- and HCV-associated HCC, which
environmental agent and a cancer that has so far been have now been delineated far enough to begin discuss‑
identified. The natural history of HCV is also variable, ing common pathways that may be tumour specific.
with up to 85% of acute infections becoming chronic The goal is to encourage the development of early bio‑
(FIG. 3). About 50% of people with chronic infections markers and specific intervention strategies that aim to
1
Department of Biology and
Sbarro Health Research (or 170 million worldwide) develop CLD, with 5–20% reduce the risk of CLD and HCC in patients who are
Organization, College of progressing to cirrhosis within 5–20 years, and 1–2% of chronically infected.
Science and Technology, these patients developing HCC per year 7. About 80% There are still many challenges ahead. With the
Temple University,
of HCC is due to HBV and/or HCV infections. There are development of rapid and sensitive screening tests, these
1900 N. 12th Street,
Philadelphia, Pennsylvania more than 250,000 new cases of HCC and an estimated viruses have been eliminated as agents of post-transfusion
19122, USA. 500,000–600,000 deaths due to this disease annually1,8. hepatitis. However, they are still widely transmitted by
2
MIT Portugal Program, The incidence of HCC is increasing, so these infections intravenous drug abuse, and to a lesser extent are also
Av. Rovisco Pais, 1049–001 remain an important public health burden. In the United sexually transmitted11,12. HBV has a highly efficacious
Lisboa, Portugal.
Correspondence to M.A.F. 
States, Europe, Egypt and Japan, more than 60% of HCC vaccine, but the implementation of vaccination pro‑
e‑mail: feitelso@temple.edu is associated with HCV, about 20% is related to HBV grammes is not uniform, and in endemic countries the
doi:10.1038/nrc3449 and chronic alcoholism contributes to the remainder. carrier frequency may still exceed 10% of the population.

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 123

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

At a glance
without virus clearance. In this context, HBV and HCV
generate proteins that blunt the development of immu‑
• Hepatitis B virus (HBV) and hepatitis C virus (HCV) are major aetiological agents of nity that should resolve chronic infections (see below).
chronic liver disease and hepatocellular carcinoma (HCC). HCC is the fifth most Although virus-specific lymphocytes are readily detect‑
prevalent tumour type and the third leading cause of cancer-related deaths able in inflammatory lesions in the liver, they are often
worldwide, which is why it is so important to find early diagnostic markers and
not sufficient to clear virus infection21. Many virus non‑
therapeutic targets, particularly when these markers and targets are common to both
chronic infections.
specific natural killer (NK) cells, NKT cells and polymor‑
• Alterations in multiple signalling pathways and patterns of host gene expression have
phonuclear leukocytes are also found in these lesions,
been documented following HBV and HCV infections, but their relative importance which mediate hepatocellular damage, but they are not
to the pathogenesis of HCC has not been clearly defined. This has limited the ability to effective in virus clearance21. Imbalances in cytokine pro‑
design appropriate therapeutic intervention strategies, and to determine the best duction may also hinder virus elimination. In chronic
time during chronic infection for their application. HBV, there seems to be a deficiency in interferon (IFN)
• The pathogenesis of HBV and HCV infections is generally immune mediated, although production, and IFN therapy in HBV-transgenic mice has
these viruses have evolved multiple mechanisms to escape immune elimination and shown strong anti-viral effects22. However, IFN reduces
to continue replicating in an infected host for many years. virus titres and disease progression in only a small num‑
• Chronic infection with either virus can result in inflammation and oxidative stress. ber of patients who are infected with HBV23 or HCV24,
A prolonged fibrotic response, resulting in cirrhosis, is also common in both infections, suggesting that success will require more than simple
which is accompanied by the appearance of localized hypoxia, rearrangement of cytokine replacement.
tissue architecture (epithelial–mesenchymal transition) and angiogenesis.
HBV- and HCV-encoded proteins alter host gene
• Altered host gene expression in chronic liver disease may be mediated by
expression and cellular phenotypes that are recog‑
epigenetic changes, by inhibition of DNA repair and/or by differential expression
nized as hallmarks of cancer. These changes promote
of microRNAs. These alterations include constitutive upregulated expression of
factors involved in ‘stemness’, suggesting that both viruses may contribute to HCC by growth factor-independent proliferation, resistance
promoting stemness. to growth inhibition, tissue invasion and metastasis,
• Early biomarkers and tumour-specific treatments for HCC are mostly lacking, although angiogenesis, reprogramming of energy metabolism, and
several signalling cascades that are activated in the liver before tumour appearance resistance to apoptosis in the face of persistent immune
suggest that oncogene addiction may be important to the pathogenesis of HCC. attack and during therapeutic intervention25. Chronic
• Understanding common mechanisms of HBV and HCV pathogenesis will help to focus inflammation also promotes genetic instability in tumour
efforts on therapeutic targets that may be most useful in the development of new cells. The contribution of HBV to HCC involves the
treatment approaches. expression of hepatitis B x (HBx) and possibly carboxy-
terminally truncated pre-S or S polypeptides (FIGS 1,4,5);
for HCV, the core protein, and non-structural (NS)
Although issues of cost, non-compliance, vaccine non- proteins NS3 and NS5A contribute to oncogenic trans‑
responders and vaccine-escape mutants have limited formation (FIGS 2,4,5). Changes in host gene expression
the effectiveness of the vaccine, universal vaccination that promote tumorigenesis also seem to support virus
is an important goal. In Taiwan, HBV vaccination has replication and/or protect virus-infected hepatocytes
Nucleoside analogues
decreased the incidence of new infections and HCC13. from immune-mediated damage or destruction. Thus,
Derivatives of standard
nucleosides that are There is no vaccine for HCV partly because infections tumour-promoting changes seem to be coincidental, as
incorporated into DNA or RNA consist of a genetically heterogeneous ‘swarm’ of virus the development of HCC does not seem to benefit these
during virus replication and particles, some of which escape neutralization14 (FIG. 3). viruses and most often kills the host.
that bring about chain Although potent antiviral nucleoside analogues against Virus persistence refers to the manner in which HBV
termination, thereby effectively
inhibiting virus replication.
the HBV and HCV polymerase, as well as protease and HCV exist in infected cells over a long period, and
inhibitors against HCV, have been developed, drug it is important to the pathogenesis of HCC. For HBV,
HBV-transgenic mice resistance is still a problem. Medium-term nucleoside virus DNA exists as an episome in the form of a mini-
Mice generated by introducing analogue treatment of chronic HBV has considerably chromosome in the nuclei of cells26. Regeneration of
a larger than full-length
reduced the incidence of HCC15. For HCV, none of the infected hepatocytes during CLD often results in the
molecular clone of hepatitis B
virus (HBV) DNA into fertilized new drugs has been in use long enough to see an effect, integration of HBx and sometimes truncated S and
ova, which are then transferred and in fact one of the most promising nucleoside ana‑ pre-S sequences into many sites in the host DNA. The
into the uterus of logues against HCV that was developed by Bristol-Myers corresponding HBx and pre-S or S polypeptides have
pseudopregnant female mice. Squibb (BMS‑986094) has been withdrawn from clinical transforming properties. Although integration into host
The offspring stably replicate
HBV from virus RNA made by
trials owing to toxicity 16. DNA may contribute to HCC through cis-acting mecha-
one or more virus DNA nisms, the large number of integration events at different
templates that are integrated Pathogenesis sites in host DNA does not support cis activation as a
into the host DNA. These mice HBV- and HCV-induced HCC develops in an environ‑ major mechanism in HCC. Instead, the products trans‑
resemble human chronic
ment of inflammation and regeneration that results from lated from the integrated templates more commonly
carriers in that they
persistently replicate virus, but chronic liver damage, suggesting that the pathogenesis of affect host gene expression by trans-acting mechanisms.
do not develop liver disease or HCC is immune mediated17 (FIG. 3). These viruses persis‑ Intracellular levels of HBx often increase after each cycle
hepatocellular carcinoma. tently replicate in cell culture without overt cellular dam‑ of hepatocellular regeneration (FIG. 4). By contrast, HCV
age or death, implying that they are noncytopathic18,19, persistence does not include integration, and maintains
Episome
A virus nucleoprotein complex
although HCV may also have cytopathic properties17. itself as an endoplasmic reticulum (ER)-associated epi‑
that replicates independently Persistent virus replication is a risk factor for HCC20 some. In infected hepatocytes, these viruses encode pro‑
of host chromatin. because inflammation often results in prolonged CLD teins that promote survival and growth, which favour

124 | FEBRUARY 2013 | VOLUME 13 www.nature.com/reviews/cancer

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

Cis-acting mechanisms the expansion of virus-infected cells following each bout reciprocal interference: when HBV markers are present,
Occur when viral DNA of hepatitis. Thus, progressive CLD seems to be asso‑ HCV markers are mostly or completely absent from
integration results in the ciated with an increase in the proportion of virus- the liver and blood. In this case, the suppressed agent
altered expression of a gene positive cells in the liver that acquire the characteristics assumes an occult infection3,28. In occult HBV infections,
adjacent to or close to the site
of integration.
of tumour cells27. the HCV core protein binds to HBx, blocking its trans‑
Persistent co‑infections in a single patient seem to activation function, and also binds to the HBV poly‑
Trans-acting mechanisms accelerate the pathogenesis of HCC through common merase and pre-genomic RNA, interfering with HBV
Occur when a virus encodes necro-inflammatory pathways or by direct oncogenic replication29. HBx in occult HBV infections stimulates
one or more proteins that
activity 28. Both exacerbate oxidative stress and the pro‑ HCV replication30. Some 30–80% of occult HBV infec‑
influence the expression of
host genes at many
duction of cytotoxic cytokines. In addition, virus muta‑ tions have integrated HBV DNA31 that expresses HBx,
chromosomal sites. tions may affect the robustness of virus replication and such patients have a lower but still increased risk of
and virus–virus interactions. These viruses also show HCC28. Among carriers with HCC, up to 23% are also
anti-HCV-positive, suggesting frequent co‑infection,
especially in countries where both viruses are endemic32.
pre-S2: 55 aa
pre-S1: 128 aa Altered innate and adaptive immunity
HBV and HCV block or alter immune responses, but
– Strand
the outcome of acute infection is influenced by the virus
inoculum, the number of hepatocytes infected and the
+ Strand S: 226 aa kinetics of the induction of innate responses. Toll-like
3,182-1 receptor (TLR) signalling, which triggers innate immunity,
inhibits virus replication in HBV-transgenic mice33.
2,800 400 However, HBV alters TLR signalling 34, resulting in
liver damage that is not effective in virus clearance.
In chimpanzees, acute HBV infection fails to trigger
innate immunity in the first weeks of infection17 (FIG. 3).
A major signalling molecule in innate immunity is nuclear
2,400
factor-κB (NF-κB), which induces pro-inflammatory
800
and hepatoprotective gene expression35. HBx activation
of NF-κB36 may suppress innate immunity by switching
NF-κB signalling to hepatoprotection, thus promoting
resistance to immune-mediated damage. Intrahepatic
NK cell function is also skewed towards cytolytic activ‑
2,000 1,200
ity without IFNγ production, suggesting that hepatocel‑
lular killing occurs without virus clearance37. Dendritic
C: 183 aa DR1 1,600 cells (DCs), which link innate immunity with adaptive
P: 832 aa
DR2 immunity, and which may be infected with HBV, are also
pre-C defective in chronic HBV infection, resulting in poor
adaptive immunity 38. CD4+CD25+FOX3P+ regulatory T
(TReg) cells help to maintain a tolerogenic environment in
the liver. These TReg cells are induced by HBx-stimulated
production of transforming growth factor‑β1 (TGFβ1)39.
In adaptive immunity, there are also defects in HBV-
X: 154 aa
specific CD8+ T cells40, which are characteristic of T cell
Figure 1 | The HBV genome and encoded genes.  Hepatitis B virus (HBV) is a small exhaustion.
DNA virus with a 3.2 kb genome. Virus particles have a partially Nature
double-stranded genome
Reviews | Cancer By contrast, HCV-infected chimpanzees show early
(indicated by a dashed line) with a cohesive overlap that spans the 5ʹ regions of each
strand and that is flanked by direct repeat sequences (DR1 and DR2). The virus encodes changes that are indicative of innate immunity 17. HCV
proteins from four genes on the minus strand of HBV DNA181. The gene S encodes the activates TLR3, which senses double-stranded RNA
major hepatitis B surface antigen (HBsAg) protein and its glycosylated partner, which are (dsRNA) in virus-replication intermediates, and retin‑
transmembrane proteins in the virus envelope. In‑frame sequences upstream from the S oic acid-inducible gene 1 (RIG1), which is an RNA heli‑
gene encode the pre-S domains, which are translated with the S sequences to make the case that recognizes the polyuridine motif of the HCV
pre-S and S polypeptides (middle and large proteins) that contain the virus receptor for 3′ untranslated region (UTR)41. Subsequent signalling
infection of hepatocytes. Gene C encodes the hepatitis B core antigen (HBcAg) which via NF-κB results in IFNβ production, which triggers
forms the nucleocapsid of the virus. The P region encodes the virus reverse transcriptase Janus kinase (JAK)–signal transducer and activator of
that also has DNA-dependent DNA polymerase activity and RNase H activity required for transcription (STAT) signalling in neighbouring cells,
virus replication. Although HBV is a DNA virus, it replicates through a pre-genomic RNA
and in the activation of IFN-target genes that promote
intermediate. Finally, the X gene encodes the small regulatory protein of the virus, the
hepatitis B x (HBx) antigen. HBx is a transactivating protein that stimulates virus gene the degradation of viral and cellular RNAs42. However,
expression and replication, protects virus-infected cells against immune-mediated HCV NS3–4A block TLR3 and RIG1 (REF. 43). In addi‑
destruction and contributes to the development of hepatocellular carcinoma. aa, amino tion, HCV core inhibits JAK–STAT44, and HCV NS5A
acid. Image is modified, with permission, from REF. 181 © (1985) Macmillan Publishers and E2 inhibit IFN signalling 45,46. Thus, innate immu‑
Ltd. All rights reserved. nity is triggered, but it is blocked by HCV (FIG. 3). With

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 125

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

a 5′ UTR 3′ UTR
III
IRES
II
Variable region

ORF
5′ Poly(U/UC) tract
I
IV

Translation and processing

b
Core E1 E2 p7 NS2 NS3 NS4A NS4B NS5A NS5B
1 192 384 747 810 1,027 1,658 1,712 1,973 2,419 3,010

Cleavage by cellular signal Autoproteolytic


Cleavage by HCV NS3
peptidase or peptidases cleavage

Figure 2 | The HCV genome and encoded proteins.  a | Hepatitis C virus (HCV) is a positive, single-stranded RNA virus
Nature Reviews | Cancer
that encodes a large polyprotein of about 3,000 amino acids from a single open reading frame (ORF). The ORF is flanked by
two untranslated regions (UTRs), which contain signals for viral protein and RNA synthesis, and for the coordination of
both processes. Translation is initiated through an internal ribosomal entry site (IRES) in the 5ʹ UTR. b | In polyprotein
maturation, the core–E1, E1–E2, E2–p7 and p7–non-structural 2 (NS2) junctions are cleaved by one or more cellular signal
peptidases to yield the structural proteins of the virus. The structural proteins of HCV consist of the core protein and the
two envelope glycoproteins (E1 and E2). p7 is an ion channel protein that promotes virus assembly. The NS2–NS3 proteins
undergo autocatalytic cleavage, which releases the NS3 serine protease. NS3 cleaves the remainder of the non-structural
polypeptides, which then assemble into membrane-associated replication complexes. NS5B (an RNA-dependent RNA
polymerase) then replicates the viral RNA. NS4B and NS5A have less well-defined roles in virus replication. NS3 also has
helicase and NTPase activities that are important for virus replication. The resulting proteins fold into a replication
complex and the mature virus buds through the membrane of the endoplasmic reticulum. The core protein, NS3 and NS5A
may also contribute to hepatocarcinogenesis. Figure is modified, with permission, from REF. 182 © (2009) American
Society for Clinical Investigation.

regard to adaptive immunity, chronic HCV infection damage, but that uninfected hepatocytes undergo
is characterized by virus-specific CD8+ T cells in the extensive apoptosis in CLD. This may explain why
liver — but they do not consistently correlate with dis‑ HBx expression correlates with the severity of CLD27.
ease severity 47 — and by weak CD4+ T cell proliferative The intra­cellular concentration of HBx increases with
responses48. Hopefully, pharmacological suppression additional integration events following each bout of
Oxidative stress
An increase in the intracellular of HCV NS3–4A and NS5B may result in the restoration of hepatitis and regeneration, resulting in outgrowth of
levels of reactive oxygen strong immune responses that aim to eliminate the virus. HBx-positive hepatocytes. For HCV, the overexpres‑
species (ROS), which are Both HBV and HCV encode proteins that have doc‑ sion of core and NS5A blocks apoptosis by activation
associated with most umented pro-apoptotic and anti-apoptotic properties, of AKT and NF‑κB, respectively 52,53. In chronic HCV,
pathological states, particularly
those involving inflammation.
depending on the experimental system used. Although FAS is upregulated in hepatocytes54, which may promote
controversial, it seems that when these viruses infect the survival of virus-infected cells by triggering apopto‑
Virus inoculum quiescent hepatocytes and promote inappropriate cell sis in inflammatory T cells. These events protect cells
The amount of virus that an growth, then the cellular response is to trigger growth from immune-mediated elimination, are pro-oncogenic
individual is exposed to during
arrest or apoptosis; however, when these viruses infect and may help to explain how these viruses persist in the
acute infection.
liver progenitor cells, or liver cells with defective nega‑ liver despite repeated bouts of hepatitis that should have
Innate immunity tive growth regulatory circuitry, growth arrest and apo‑ eliminated the virus.
Comprises responses that ptosis do not occur 49,50. Whether apoptosis occurs may
develop rapidly, but that are also depend on the intracellular concentration of the Oxidative stress
antigen nonspecific. Often
mediated by molecules binding
virus oncoproteins. Among infected hepatocytes with HBx and the HCV core protein are associated with mito‑
to toll-like receptors. low levels of virus oncoproteins, cell cycle arrest and chondria55,56 where they trigger oxidative stress that may
apoptosis predominate, but among cells with progres‑ result in apoptosis57,58 (FIG. 4). Proteomic and metabolite
Adaptive immunity sively higher intracellular levels of virus oncoproteins, profiles from HBV-transgenic mice have shown evi‑
Comprises responses that are
increasingly strong survival and proliferative signals dence of oxidative stress59. In these mice, oxidative stress
primed by innate immunity
and that are strong and antigen override negative growth regulation, resulting in resist‑ triggered an antioxidative response, exemplified by the
specific. ance to apoptosis. Resistance to apoptosis not only pro‑ activation of forkhead box O4 (FOXO4), which confers
motes virus persistence, but it is also characteristic of resistance to oxidative cell death60. In HBV replication,
Tolerogenic environment tumour cells (FIG 4). For HBV, sustained high levels of HBx promotes cytosolic calcium signalling, resulting in
An environment in which
immune responses are not
HBx block tumour necrosis factor-α (TNFα)- and FAS- the accumulation of Ca2+ in mitochondria and increased
readily triggered against mediated apoptosis by activation of NF‑κB51, suggesting levels of reactive oxygen species (ROS)61, as well as the
antigens. that infected hepatocytes survive immune-mediated activation of the PYK2 and SRC kinases, which both

126 | FEBRUARY 2013 | VOLUME 13 www.nature.com/reviews/cancer

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

Acute, resolving infection the ER, followed by uptake into mitochondria, result‑
ing in decreased mitochondrial permeability transition,
and the generation of ROS65. In both cases, activation
Rapid, strong, multi-specific
immune responses of ROS-sensitive transcription factors stimulates virus
25% 15% replication and cell growth, and contributes to tumour
development (FIG. 4).
HBV HCV When ER stress occurs, protein-folding capacity is
10% 85% compromised and the unfolded protein response (UPR)
Fails to trigger Blocks innate
innate immunity immunity is activated to correct the protein defect and to restore ER
function. Both HBV (HBs and HBx) and HCV (NS3–4A)
alter Ca2+ signalling and increase ROS levels, which trigger
Slow, weak adaptive immunity
HBx pre-S and S to few virus antigens resulting NS3 NS5A Core ER stress and the UPR66,67 (FIG. 4). However, when ER stress
in virus persistence becomes chronic during virus infection, UPR signalling
cannot be sustained, and autophagy is activated to help to
70% 50%
degrade the unfolded proteins and to maintain cell viabil‑
ity. Autophagy is triggered by both viruses67,68 in order to
Asymptomatic Asymptomatic help restore ER integrity, which is important for viral rep‑
carriers 30% 50% chronic infections
lication67,69. Autophagy also promotes cell survival70 and
virus persistence, which are risk factors for HCC71 (FIG. 4).
May resolve Hepatitis
Hypoxia and angiogenesis
May regress or
In chronically infected livers, the development of cir‑
Fibrosis
become inactive rhotic nodules is accompanied by a progressive decrease
of vasculature. This results in localized regions of reduced
Cirrhosis oxygen tension, and the development of a hypoxic
environment (FIG. 4). The cellular response to hypoxia
Promotes hallmark changes of cancer involves upregulated expression of HIF1α, which pro‑
motes angiogenesis by transcriptionally upregulating
vascular endothelial growth factor (VEGF), elevates
HCC
pro-inflammatory prostaglandin levels by stimulating
Figure 3 | Immunity in the natural history of HBV and HCV infections.  Immunity has cyclooxygenase 2 (COX2) and enhances cell migration
an important role in the outcome of acute infections. Rapid, strong Nature Reviews | Cancer
and multi-specific by activation of matrix metallo­proteinases (MMPs)72.
responses against many hepatitis B virus (HBV) and hepatitis C virus (HCV) proteins often In HBV infection, HBx binds to and stabilizes HIF1α
result in an acute infection that is cleared by the host. The failure of HBV to trigger innate
and stimulates HIF1α transcription73,74, thus promot‑
immunity results in suboptimal priming of adaptive immune responses that are crucial to
ing angiogenesis and cell ‘stemness’ (REF. 75) (see below).
virus clearance, and this promotes the onset of chronic infections. Before the HBV vaccine,
this was most likely to occur in infants and young children with immature immune HBx also promotes angiogenesis by stimulating the ERK
systems. Approximately 65% of acute HBV infections of adults are subclinical; about 1% of MAPK, which upregulates the pro-angiogenic growth
acute adult infections result in fulminant hepatitis, which is rapid, highly aggressive and factor angiopoietin 2 (ANG2)76 (FIG. 4). HCV core, E1,
may be fatal (not shown). For HCV, innate immunity is triggered on acute infection, but is NS3 and NS5A also upregulate HIF1α under hypoxic
blocked by the virus, again resulting in poor adaptive immunity. In addition, acute HCV conditions, resulting in increased VEGF, COX2, ANG2
infections consist of a ‘swarm’ or ‘quasi-species’ in which different virus particles consist of and several MMPs72,76,77 (FIG. 4). Hypoxia also aggravates
slightly different sequences in their neutralizing determinants. The most prevalent of cell damage and inflammation, inhibits liver regenera‑
these quasi-species trigger neutralizing antibodies, but by the time this happens, minor tion and is a major stimulus in the pathogenesis of HCC.
quasi-species become dominant and new quasi-species arise owing to the lack of
The importance of angiogenesis has been highlighted by
proofreading by the virus-encoded polymerase during virus replication. The result is that
the use of the multikinase inhibitor sorafenib10, which
the virus is often one step ahead of the ability of the host to develop neutralizing
antibodies, and this often contributes to the appearance of chronic infections. HBx, targets the VEGF receptor (VEGFR) and other signal‑
hepatitis B x; HCC, hepatocellular carcinoma; NS, non-structural. ling pathways. However, it has yielded only transitory
improvements followed by the appearance of resistance.
HIF is also stabilized by increased insulin-like growth
promote HBV replication and participate in early steps factor receptor 1 (IGFR1), epidermal growth factor recep‑
of HCC62 (FIG. 4). HBx in mitochondria binds to the tor (EGFR) and PI3K–AKT signalling. Both EGFR and
HIF1 voltage-dependent anion-selective channel protein 3 PI3K–AKT signalling are activated by HBx and HCV
A transcription factor (VDAC3), leading to membrane depolarization and to (FIG. 4). Increased EGFR signalling seems to mediate the
comprised of an α-subunit and
the production of additional ROS56. transition from cirrhosis to HCC in rats78. EGFR signals
a β-subunit that triggers
changes in gene expression HCV core protein-transgenic mice that developed through heterodimerization with ERBB2. ERBB2 is upreg‑
that promote cell survival HCC had impaired function of the mitochondrial res‑ ulated and co‑stains with HBx in the livers of patients with
under hypoxic conditions. piratory chain, increased ROS levels and decreased HCC, suggesting ERBB2 activation79 (FIG. 4). Increased
function of several antioxidants63. Activation of hypoxia- ERBB2 levels are also present in the livers of chronic
Autophagy
An adaptive response that
inducible factor 1 (HIF1) resulted in the upregulation of HCV-infected and tumour-bearing patients80. These
promotes cell survival under target genes, including those for glycolytic enzymes64. pathways of HIF activation underscore the importance
conditions of stress. In the HCV-infected cell, there is a release of Ca2+ from of hypoxia in the pathogenesis of HCC (FIG. 4).

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 127

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

Senescence Oncogenes, tumour suppressors and senescence and prognosis of early cancer 81. Oncogene-induced
Occurs when a particular cell Altered expression of senescence-related genes is often senescence has been linked to two signalling pathways
type has undergone a associated with the activation of RAS signalling 81, that are often disrupted in cancer: AKT–ARF–p53–p21
predetermined number of cell which is activated by HBx 82. HBx is also capable of (also known as WAF1) and RAS–MEK–ERK–INK4A
divisions. These cells remain
metabolically active but no
overcoming RAS-induced senescence in immortal‑ (also known as p16)–RB, both of which operate in
longer divide. ized cells83. Activation of RAS in pre-malignant tissues HCC84 and are activated by HBx 85,86. Thus, overcom‑
showed that senescence exists before tumour appear‑ ing senescence seems to be a major step in early tumour
ance, suggesting that it may be valuable in the diagnosis development.

a
Low HBx Increased HBV High HBx
levels in liver DNA integration levels in liver

Regeneration Changes in liver architecture

Inflammation Hepatitis Fibrosis Cirrhosis HCC Metastasis

b TGFβ
TGFβR IGFR1 EGFR ERBB2 E-cadherin

RAS–ERK
β-catenin

PI3K–AKT HBx HCV EMT


proteins

Cirrhosis
• RAS–ERK
SMADs to HCC SRC Increased HIF,
• PI3K–AKT Resistance
transition hypoxia and
to apoptosis
angiogenesis

HCV HBx
proteins Stemness VEGF, COX2,
Mitochondrial (NANOG, KLF4, ANG2, MMPs
stress OCT4 and MYC) and NOTCH

HCV
proteins Mitochondria MYC, MDR1
ROS and TERT
HBx
ER
stress
HDAC HBx
Autophagy UPR SMADs
A2M
NF-κB NF-κB
Increased cell FN
SNAIL CDH1
survival and HCC risk
persistent virus factors EMT
replication β-catenin EPCAM
TGFB1
ER
Nucleus

TGFβ

Figure 4 | Signalling pathways targeted by HBV and HCV in CLD pathogenesis. a | Major steps in the
Nature pathogenesis
Reviews | Cancer
of hepatitis B virus (HBV) and hepatitis C virus (HCV)-associated hepatocellular carcinoma (HCC) are shown. In both virus
infections, immune responses to persistent virus replication promote the development of hepatitis, which consists of
hepatocellular destruction followed by regeneration and/or fibrosis. During regeneration, the hepatitis B x (HBx) and
pre-S and S genes increasingly integrate into host DNA, resulting in increased levels of intracellular HBx. b | The activities
and expression of selected signalling pathways and target genes are altered by HBV- and HCV-encoded proteins. Note
that both viruses promote the growth and survival of infected cells during chronic liver disease (CLD), and activate several
overlapping signalling pathways (for example, RAS, PI3K, epidermal growth factor receptor (EGFR) and insulin-like growth
factor receptor 1 (IGFR1)) that contribute to the pathogenesis of HCC. A2M, α2‑macroglobulin; ANG2, angiopoietin 2;
CDH1, E-cadherin; COX2, cyclooxygenase 2; EPCAM, epithelial cell adhesion molecule; EMT, epithelial–mesenchymal
transition; ER, endoplasmic reticulum; FN, fibronectin; HDAC, histone deacetylase; HIF, hypoxia-inducible factor; TERT,
human telomerase reverse transcriptase; MDR1, multi-drug resistance 1; MMP, matrix metalloproteinase; NF-κB, nuclear
factor-κB; ROS, reactive oxygen species; TGFβ, transforming growth factor-β; UPR, unfolded protein response; VEGF,
vascular endothelial growth factor.

128 | FEBRUARY 2013 | VOLUME 13 www.nature.com/reviews/cancer

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

Inflammation, oxidative and oncogenic stress in These viruses overcome senescence by inhibiting the
chronic HBV and HCV infections accelerate cellu‑ expression and function of tumour suppressors (FIG. 5).
lar senescence. Senescence limits the proliferation of Both viruses upregulate DNA methyltransferases
damaged cells and reduces the risk of malignancy 87 (DNMTs)89,90, which decrease the expression of genes
by triggering the expression of tumour suppressors88. involved in DNA repair 91. HBx, as well as the HCV
core, NS3 and NS5A proteins, compromise several
a p53 functions53,92, including nucleotide excision repair
HBx β-catenin miR-181 TIMP3 and transcription-coupled repair 93. HBx also promotes
the methylation of the ankyrin-repeat-containing,
• MMP2 SH3‑domain-containing and proline-rich-region-
Tumorigenicity
• MMP9
containing proteins (ASPP1 and ASPP2) that activate
EPCAM Stemness p53 (REF. 94). HCV also blocks DNA repair through the
ataxia telangiectasia mutated (ATM) DNA repair path‑
miR-148a PTEN PI3K–AKT–mTOR Loss of tumour way 95. In addition, HCV core protein and HBx suppress
suppressors, DNA the cyclin-dependent kinase (CDK) inhibitors INK4A
repair and genome and p21 via promoter methylation, resulting in the
miR-373 CDH1 EMT integrity
inactivation of the RB tumour suppressor 96,97 (FIG. 5).
p53 Transcription- A microRNA (miRNA), miR‑221, which is upregulated
ERCC3
coupled repair in HBV- and HCV-related HCC, blocks the expression
of the CDK inhibitor p27 and promotes tumour growth
HCV miR-141 DLC1 and progression by activation of the PI3K–AKT–mTOR
proteins
pathway 98. Thus, abrogating the pathways that are
populated by tumour suppressors and CDK inhibitors
b HBx TFIIH p53 compromises the ability of cells to mediate DNA repair,
DNA excision repair thereby promoting genetic instability 99.
HCV
proteins DNMT ATM The ability of HBV and HCV to alter the expres‑
sion of selected miRNAs affects virus persistence,
miR-602
as well as oncogene and tumour suppressor path‑
RASSF1A ways in HCC (FIG. 5). For example, miR‑122, which
Loss of tumour
suppressors, DNA promotes hepatocellular differentiation and tumour
repair and genome suppression, blocks cyclin G1 expression and inhibits
miR-152 DNMT1 INK4A RB integrity HBV replication by a miR‑122–cyclin G1–p53–HBV
CDKN1A
enhancer regulatory pathway 100. HBV increases miR‑122
expression, which targets HBx sequences101, thereby
ASPP1 and ASPP2 p53 promoting virus persistence. For HCV, miR‑122
stimulates virus replication by stabilizing the virus
CDH1 EMT RNA in infected liver 102, and is suppressed in HCC100,
which may explain why HCV replication is generally
GSTP1 ROS AP1 miR-21
lower in HCC compared with adjacent liver. However,
miRNA arrays using virus-infected compared with
uninfected liver showed that pathways that are related
Inflammation
miR-122 Virus persistence and metastasis
to cell death, DNA damage, recombination and sig‑
nal transduction were activated in HBV-infected liver,
Figure 5 | Examples of epigenetic alterations in host gene expression that are and those related to immune response, antigen pres‑
relevant to the pathogenesis of HBV- and HCV-associated HCC. Part a shows a few entation, proteasome, cell cycle and lipid metabolism
of the pathways that distinguish the mechanisms through which hepatitis
Nature B virus
Reviews (HBV)
| Cancer
were activated in HCV-infected liver 103. Although
and hepatitis C virus (HCV) contribute to hepatocarcinogenesis, and part b illustrates
some of the shared pathways. Hepatitis B x (HBx)-upregulated expression of the these data suggest differences in the mechanisms
microRNA (miRNA) miR‑181 promotes tumorigenicity, and upregulated expression of of HCC used by these viruses, many miRNAs target
miR‑148a and miR‑602 blocks the function of crucial tumour suppressor proteins. HBV overlapping signalling pathways and genes, suggesting
and HCV upregulation of miR‑21 in an environment of oxidative stress promotes that further analysis may reveal common pathways
inflammation in the liver and metastases during tumour progression. Both viruses also of oncogene activation and inactivation of tumour
affect p53 and RB function, resulting in the loss of genome integrity and uncontrolled suppression.
cell cycle progression, respectively. Thus, inactivation of tumour suppressors and the
activation of oncogenic pathways, some of which are regulated via virus alterations in EMT and fibrogenesis
miRNA expression profiles, seem to be important to hepatocarcinogenesis. AP1, Epithelial–mesenchymal transition (EMT) is important
activating protein 1; ATM, ataxia-telangiectasia mutated; CDH1, E-cadherin; CDKN1A,
in the pathogenesis of CLD as it promotes fibrogenesis,
cyclin-dependent kinase inhibitor 1 (encodes p21); DLC1, deleted in liver cancer 1;
DNMT1, DNA methyltransferase 1; EPCAM, epithelial cell adhesion molecule; ERCC3, tumour progression and metastasis. In fibrogenesis, type 2
excision repair cross-complementing rodent repair deficiency, complementation EMT occurs in response to continuing inflammation,
group 3; GSTP1, glutathione S‑transferase P1; MMP, matrix metalloproteinase; ROS, ultimately leading to organ destruction. De novo expres‑
reactive oxygen species; TFIIH, transcription factor II H; TIMP, tissue inhibitor of sion of mesenchymal proteins104 — such as fibroblast-
metalloproteinase 3. specific protein 1 (FSP1; also known as S100A4),

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 129

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

Epithelial–mesenchymal collagen I and collagen III, and α‑smooth muscle Do these viruses promote stemness?
transition (α‑SMA) — and upregulated expression of base‑ Both viruses promote characteristics of cancer stem
(EMT). A transient and ment membrane-degrading MMP2 and MMP9 by cells (CSCs) (FIG. 4), which have an important role in
reversible switch from a hepatic stellate cells (HSCs)105 that transdifferenti‑ the pathogenesis of HCC126–129. Stemness markers (such
polarized, epithelial phenotype
to a fibroblastoid or
ate into myofibroblasts 106 (and possibly hepatocytes as NANOG, OCT4, SOX2 and Krüppel-like factor 4
mesenchymal phenotype. EMT having undergone EMT)107 retain a permanent mes‑ (KLF4)) are reactivated130,131 and expressed in HCC132,
is divided into three categories: enchymal state that contributes to fibrosis108. TGFβ, as well as in many other tumour types 133. HBx pro‑
type 1 occurs in development which promotes EMT by regulating transcription motes the expression of NANOG, KLF4, OCT4 and
(associated with embryo
factors that suppress epithelial markers (for example, MYC, as well as the stemness-associated markers epi-
implantation, formation and
organ development); type 2
components of cell junctions) and which enhances thelial cell adhesion molecule (EpCAM) and β‑catenin132.
occurs in fibrosis (associated mesenchymal features, is central to the generation Although normal adult hepatocytes do not express
with tissue damage and of myofibroblasts. Once activated, myofibroblasts EpCAM, HBV-infected hepatocytes do express this
inflammation, and generates secrete TGFβ and sustain their own activation by protein, implying that HBx confers stemness properties
repair-associated
mesenchymal cells and
an autocrine mechanism. TGFβ1 promotes the tran‑ to at least some infected cells132. NANOG and MYC,
fibroblasts); and type 3 occurs scription of genes encoding collagen I and collagen as well as the stemness-associated markers β‑catenin,
in cancer and metastasis. III 109, as well as SNAIL 110, which transcriptionally CD‑133 and LIN28, were also upregulated by HCV134.
suppresses E‑cadherin 111 (FIG.  4) . HBx upregulates Stabilization of β‑catenin promotes the transcription
Fibrogenesis
TGFβ1 (REF. 39) by SMAD-dependent (via stabili‑ of stemness genes in association with TCF/LEF1,
Involves the synthesis and
deposition of extracellular
zation of the SMAD4 complex) 112 and non-SMAD- OCT4 and NANOG135. β‑catenin also transcriptionally
matrix proteins, and is a form dependent pathways (via activation of RAS–ERK upregulates EpCAM132. EpCAM, in turn, contributes
of tissue repair that follows a and PI3K–AKT)113. HCV core protein activation of to stemness through its binding to OCT4, NANOG,
bout of hepatitis, in which TGFβ34 also promotes EMT in hepatocytes and HCC SOX2 and KLF4 promoter regions136. In HCC xeno‑
hepatocytes are injured and
destroyed.
cells. HBx and HCV core also seem to convert TGFβ1 graft experiments, only EpCAM+ cells efficiently gener‑
signalling from negative to positive growth regulation ated invasive tumours, suggesting that EpCAM+ cells
Cancer stem cells and shift TGFβ responses from tumour suppression display CSC-like properties such as tumour initiation
(CSCs). A subset of tumour to EMT112,114. and progression137. Both HBV and HCV promote the
cells that share properties with
Liver fibrogenesis (type 2 EMT) and HCC metasta‑ expression of one or more miR‑181 family members,
normal tissue stem cells,
including self-renewal (by
sis (type 3 EMT) are also mediated by miR‑21, which which are known to upregulate EpCAM129,138 (FIG. 5).
symmetric and asymmetric is upregulated by NF-κB in HBV- and HCV-associated miR‑181 members are highly expressed in embry‑
division) and the capacity to HCC115. Activated TGFβ and RAS–ERK–AP1 result onic livers, in HSC, and in patients with α-fetoprotein
differentiate. in the inhibition of multiple tumour suppressors116. (AFP)-positive tumours, suggesting that they help to
Epithelial cell adhesion
Importantly, these signalling pathways are also acti‑ maintain stemness138.
molecule vated by HBx and HCV 36,53,80,82,86,112,114. As miR‑21 Stemness is favoured by hypoxia, which enhances
(EpCAM). A cell adhesion activation usually occurs during CLD, in which ROS the activity of OCT4, MYC and NANOG139. The effect
molecule found on hepatic stimulates NF-κB and AP1, these observations provide of hypoxia on CSCs is primarily mediated by HIFs.
stem cells and cancer stem
a link between chronic inflammation and early events Although the detailed mechanisms that link hypoxia to
cells; when activated by
proteolysis, an intracellular
in HCC. stemness are not known, β‑catenin can enhance the tran‑
domain reaches the nucleus, SRC signalling, which is activated by both viruses, scription of HIF1A, thereby promoting cell survival and
where it triggers cell cycle is important to EMT 117,118 (FIG. 4). SRC is known to adaptation to hypoxia140. HIF upregulates MYC, which
progression and mitosis by promote the expression of MMPs by multiple mecha‑ transactivates human telomerase reverse transcriptase
upregulating cyclin E and MYC
expression.
nisms, including activation of the ERK and PI3K (TERT)141, thus linking two key regulators of stem cell
pathways. HBx-stimulated SRC promotes EMT119 by biology and cancer. Hypoxia also stimulates EMT in
Telomerase reverse destabilization of adherens junctions120. This is asso‑ HCC142 and in hepatocytes, suggesting that HIF1α
transcriptase ciated with tyrosine phosphorylation of β‑catenin (which is upregulated by both viruses)73,74,77 (FIG. 4) is
(TERT). Helps to prevent the
(activation) in E‑cadherin complexes 120. In addi‑ important for EMT143. In other tumour types, EMT pro‑
erosion of chromosomal ends
(telomeres) after each cycle of
tion, E‑cadherin is suppressed by the HBx-mediated vides an environment that promotes the appearance of
cell division. Constitutive upregulation of SNAIL121, and HCV (NS5A) upregu‑ CSCs144 that may facilitate escape to a new niche. EMT
overexpression of telomerase lation of TWIST2 (REF. 122). E‑cadherin expression and stemness share several signalling pathways that are
(of which TERT is a component) is also suppressed by promoter DNA methylation, constitutively activated by these viruses (FIG. 4). In ovarian
is characteristic of many
tumour types.
which is mediated by HBx 121 and the HCV core pro‑ cancer, the acquisition of stemness ensures the generation
tein 123 (FIG. 5). Further, these viruses promote EMT of the crucial tumour mass required for progression and
by activation of RAS, WNT and VEGF signalling, metastases145. This may also occur in HCC. The finding
which all activate SNAIL. HCV core and E2 have that EMT generates self-renewing stem cells provides
also been implicated in fibrogenesis by triggering a link between tumour metastasis and CSCs. In other
EMT 124,125. Thus, activated SRC signalling by these words, cancer cells that undergo an EMT are capable of
viruses contributes to early steps in HCC. During metastasizing through their acquired invasiveness and
natural infection, fibrosis may resolve spontaneously then initiate metastases through self-renewal potential.
or may resolve as a result of sustained inhibition of Detailed understanding of the origin and molecular pro‑
virus replication, but generally, advanced fibrosis and file of liver CSCs, and how normal stem cells in the liver
cirrhosis may become inactive (that is, arrested) but differ from CSCs, may allow them to be targeted, thus
not resolved. improving therapeutic outcome.

130 | FEBRUARY 2013 | VOLUME 13 www.nature.com/reviews/cancer

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

Biomarkers and therapies Des-γ-carboxy prothrombin (DCP) and Lens culi‑


A major difficulty in developing early biomarkers naris agglutinin-reactive fraction of AFP (AFP‑L3)
of HCC comes from an incomplete understanding of have a higher specificity and a better negative pre‑
tumour pathogenesis, and the extent of marker variability dictive value compared with total AFP. This combi‑
among patients. This is why personalized therapy is not nation has identified patients with negative imaging
yet a reality for patients with HCC. In one study, a sub‑ results for tumours who would benefit from follow‑up
set of patients with HCC who had elevated intrahepatic evaluation152. Recently, glypican‑3 (GPC3), which is a
EpCAM and β‑catenin levels had a poor prognosis relative membrane-bound heparin sulphate proteoglycan, was
to EpCAM-negative patients with HCC146. Independent found to be expressed in HCC cells and tissues, and to
work has shown a considerable difference in the survival some extent in hepatocytes. Although GPC3 seems
of patients with HCC post-diagnosis based on elevated to be an early marker of HCC153, it is not clear whether
expression of ERBB2 (REF. 79). Both EpCAM and ERBB2 it is virus associated. Serum GPC3 levels were
are functionally connected to β‑catenin, which promotes increased in most patients with early-stage HCC, and
stemness at early stages of HCC. If further work provides an assay combining GPC3 with AFP has increased
validation that any of these markers have prognostic sensitivity for the detection of HCC at all stages. Other
value, their detection before tumour appearance may also potential serum biomarkers for HCC include heat
suggest their usefulness as therapeutic targets. shock protein 27 (HSP27), α2‑HS glycoprotein, apoli‑
Global hypomethylation of DNA also alters gene poprotein E and apolipoprotein M, and several autoan‑
expression before tumour appearance, which turns on tibodies154.Unfortunately, most of these have not been
genes that promote growth, as does selected hyper‑ validated in larger studies, and given that HCC is
methylation of tumour suppressor genes, including heterogeneous in terms of markers, it is not clear
IGF-binding protein 3 (IGFBP3), RASSF1A, INK4A, which marker or markers are going to be useful for
E‑cadherin and glutathione S‑transferase P1 (GSTP1)147 early detection.
(FIG. 4). In chronic HBV infection, HBx suppression of Additional attempts to identify biomarkers have used
miR‑152 upregulates DNMT1, which methylates the microarrays of tumour compared with non-tumour or
promoters of most of these tumour suppressors92,147. uninfected liver and cirrhotic liver compared with livers
DNA methylation signatures may help to classify with dysplastic and/or regenerative nodules. Few con‑
patients for treatment purposes, but their clinical sistent differences have been found155–160, and these
usefulness remains to be fully evaluated. differences have not been validated by larger studies.
The expression of selected miRNAs is profoundly Differences might also be due to the techniques used,
altered in many cancers, including HCC, suggesting their prior treatment, stage of tumour differentiation, encap‑
potential importance as biomarkers and therapeutic tar‑ sulation, vascularization, extent of fibrosis or inflamma‑
gets (FIG. 5). In addition to patient-to‑patient and disease tion and whether the analysed samples were from primary
variability, these analyses use different approaches and tumours or intrahepatic metastases. Further, it is not clear
starting materials for analyses (HCC versus non-tumour whether these studies identify markers of early HCC or
liver or HCC compared with uninfected liver). Although of tumour progression. Biomarkers have also been sug‑
the mechanisms by which HBV and HCV alter miRNA gested from studies on cell lines and from animal mod‑
expression remain to be elucidated, many miRNA genes els of HBV and HCV, but there has been little large-scale
are located at fragile sites or in cancer-associated genomic validation using human clinical samples. However, recent
regions, suggesting that the genetic instability that is work has identified a single nucleotide polymorphism
associated with these chronic infections may alter the (SNP) in EGF that is associated with an increased risk of
expression of these miRNAs148. The potential of miRNAs HCC among patients with HCV161,162. An SNP in inter‑
as biomarkers derives from an appreciation that most leukin‑28B (IL28B) also influences the outcome of HCV
miRNAs target many genes, and that as cancer is a infection163 and the response of patients to treatment 164.
multi-step process, altered expression of crucial miRNAs By contrast, it is not clear that SNPs in EGF and IL28B
will probably have diagnostic or prognostic value. have any role in the pathogenesis of HBV infection165–167.
Selected miRNAs may also be therapeutic targets, such as By comparison, SNPs in TGFB1 may increase the risk of
miR‑122, which stimulates HCV replication. Currently, HCC in both infections168,169. Thus, there is a potential role
a miR‑122 antagonist (Miravirsen (SPC3649; Santaris for SNPs as biomarkers.
Pharma A/S)) is being evaluated in patients with chronic HCC is not detected early enough for curative inter‑
HCV, and preliminary results have shown considerable vention. Available therapies are not tumour specific
and prolonged anti-viral activity (see the ClinicalTrials. and are not very effective when HCCs are diagnosed
gov website; see Further information). This approach at a late stage. The first drug found to increase sur‑
holds promise for the treatment of HCV, and it is hoped vival, sorafenib, does so by inhibiting cell proliferation
that crucial pathways in tumorigenesis could be similarly (via the RAS–RAF–ERK pathway) and by blocking
targeted. However, as miR‑122 can be a tumour suppres‑ angiogenesis (via the VEGFR and platelet-derived
sor, blocking its expression may increase the risk of HCC growth factor receptor). HCC is also associated with
in patients who are chronically infected149,150. the activation of EGFR, IGFR, WNT–β‑catenin and the
In serum, elevated AFP >400 ng per ml has been PI3K–AKT–mTOR pathways 51,53,170–175, which may
used as an early marker of HCC151, although it lacks underlie sorafenib resistance, suggesting that combined
sensitivity and is not specific for virus-associated HCC. therapies will be required to have a sustained effect.

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 131

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

There are more than 50 drugs in almost 200 clinical This provides opportunities for the development of spe‑
trials that are currently underway or that have recently cific immunotherapeutics and/or therapeutic vaccines
been completed in patients with HCC176 but it is not that would help to resolve virus infection and block
yet clear what pathways support HCC growth. In this CLD progression. Immunotherapy may also be com‑
context, it will be important to target pathways that are bined with nucleoside analogues and other inhibitors
involved in oncogene addiction, as these drive tumour of virus replication in the control of these infections.
growth. HBV-associated HCC seems to be addicted A cure may be possible for HCV because it exists solely
to the RAS–ERK and PI3K–AKT–mTOR pathways87. as an episome, but for HBV, proteins made from inte‑
However, it is not clear what pathways define onco‑ grated virus templates may still trigger and promote
gene addiction in HCV-mediated HCC. The transcrip‑ CLD even if episomal DNA (which includes linear
tion factor LSF, which promotes cell cycle progression replicative forms and covalently closed circular DNA
through G1/S phase, may be part of another onco‑ (cccDNA)) is cleared. The identification of signalling
gene addiction pathway in HCC177. WNT–β‑catenin pathways and gene targets that are common to the
and ERBB2 signalling are constitutively activated in pathogenesis of HBV and HCV infections may provide
HCC79,87, but it is not clear whether HCC is addicted opportunities for the application of existing drugs and
to these pathways. Given the heterogeneity of HCC, the development of new drugs that will be effective
treatment cocktails will need to be developed for dif‑ against both infections. Further, it is likely that simulta‑
ferent subsets of patients depending on their bio‑ neously targeting multiple pathways that drive tumour
marker profiles. As clinical trials cannot be conducted progression will provide opportunities to develop spe‑
over many years or decades, biomarkers may be used cific therapies. The importance of developing treat‑
as surrogate end points in evaluating drug efficacy. ments for patients who are chronically infected before
This should accelerate the evaluation of new drugs in tumour appearance is underscored by the reality that
patients with CLD who are at a high risk of HCC before most of the changes in the liver before tumour develop‑
tumour appearance, and may allow the development of ment are reversible by drugs, but many of the changes
preventive chemotherapeutics. in tumour nodules involve mutations in pathways that
are much more difficult to treat.
Conclusions Importantly, virus proteins trigger changes in sig‑
HBV and HCV trigger changes in gene expression in nalling pathways and gene expression in hepatocytes
the liver that confer the hallmarks of cancer on cells. that are often found to be mutated in HCC178. In such
These changes result from alterations in DNA methy­ a model, which has been proposed for other tumour
lation, changes in miRNA expression profiles and the types, epigenetic modification of tumour suppressor
constitutive activation of numerous signal transduc‑ genes may permit the constitutive expression of onco‑
tion pathways. The induction of EMT in cirrhotic or genes in early tumorigenesis, and mutation in these
dysplastic and early HCC nodules is associated with a same oncogenes results in higher constitutive expres‑
hypoxic environment that favours the survival and out‑ sion that is characteristic of tumour progression179.
growth of cells with stem cell-like characteristics that Thus, oncogene addiction may be an early event in
promote tumour development and progression. The cancer. Inflammation and ROS would promote this
pathogenesis of HCC is immune mediated, and the addiction. As cell clones expand, they acquire her‑
oxidative environment created in the liver by immune itable epigenetic changes that result in a permanent
responses is exploited by these viruses to constitutively change in phenotype179, and take on increasing fea‑
activate signalling pathways that block apoptosis, pro‑ tures of tumour cells. These changes will be fertile
mote survival, support HBV and HCV replication, ground for the establishment of biomarkers and
and promote virus persistence. These viruses also targets for therapeutic intervention. Fortunately,
delay or block innate immunity, thereby compromis‑ drugs are already being developed to many of these
ing the development of adaptive immune responses. epigenetic targets180.

1. El‑Serag, H. B. & Rudolph, K. L. Hepatocellular 7. Strader, D. B. et al. Diagnosis, management, and 12. Smyth, R., Keenan, E., Dorman, A. &
carcinoma: epidemiology and molecular carcinogenesis. treatment of hepatitis C. Hepatology 39, 1147–1171 O’Connor, J. Hepatitis C infection among
Gastroenterology 132, 2557–2576 (2007). (2004). injecting drug users attending the National Drug
2. Beasley, R. P., Hwang, L. Y., Lin, C. C. & 8. Venook, A. P., Papandreou, C., Furuse, J. & de Treatment Center. Ir. J. Med. Sci. 164, 267–268
Chien, C. S. Hepatocellular carcinoma & HBV. A Guevara, L. L. The incidence and epidemiology of (1995).
prospective study of 22,707 men in Taiwan. Lancet 2, hepatocellular carcinoma: a global and regional 13. Ni, Y. H. & Chen, D. S. Hepatitis B vaccination in
1129–1133 (1981). perspective. Oncologist 15, S5–S13 (2010). children: the Taiwan experience. Pathol. Biol. (Paris)
3. Lee, W. M. Hepatitis B virus infection. N. Engl. J. Med. 9. Poon, R. T. et al. Improving survival results after 58, 296–300 (2010).
337, 1733–1745 (1977). resection of hepatocellular carcinoma: a prospective This body of work showed that immunization
4. Roberts, L. R. & Gores, G. J. Hepatocellular study of 377 patients over 10 years. Ann. Surg. 234, against HBV also prevented the development of
carcinoma: molecular pathways and new 63–70 (2001). liver cancer, making it the first cancer vaccine.
therapeutic targets. Semin. Liver Dis. 25, 212–225 10. Llovet, J. M. et al. Sorafenib in advanced 14. Pawlotsky, J. M. Hepatitis C virus population dynamics
(2005). hepatocellular carcinoma. N. Engl. J. Med. 359, during infection. Curr. Top. Microbiol. Immunol. 299,
5. Sherman, M. Risk of hepatocellular carcinoma in 378–390 (2008). 261–284 (2006).
hepatitis B and prevention through treatment. Cleve. The development and application of sorafenib 15. Papatheodoridis, G. V., Lampertico, P.,
Clin. J. Med. 76, S6–S9 (2009). represents the first time that targeting molecular Manolakopoulos, S. & Lok, A. Incidence of
6. Nguyen, V. T., Law, M. G. & Dore, G. J. Hepatitis pathways known to operate in HCC has resulted in hepatocellular carcinoma in chronic hepatitis B
B‑related hepatocellular carcinoma: epidemiological an increase in patient survival. patients receiving nucleos(t)ide therapy: a
characteristics and disease burden. J. Viral Hepat. 16, 11. Maddrey, W. C. Hepatitis B, an important public systematic review. J. Hepatol. 53, 348–356
453–463 (2009). health issue. J. Med. Virol. 61, 362–366 (2000). (2010).

132 | FEBRUARY 2013 | VOLUME 13 www.nature.com/reviews/cancer

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

16. Cohen, J. Infectious disease. Despite setbacks, 41. Saito, T., Owen, D. M., Jiang, F., Marcotrigiano, J. & 63. Fujinaga, H., Tsutsumi, T., Yotsuyanagi, H., Moriya, K.
optimism on drugs for hepatitis C. Science. 337, Gale, M. Jr. Innate immunity induced by composition- & Koike, K. Hepatocarcinogenesis in hepatitis C: HCV
1450–1451 (2012). dependent RIG‑1 recognition of hepatitis C virus RNA. shrewdly exacerbates oxidative stress by modulating
17. Rehermann, B. & Nascimbeni, M. Immunology of Nature 454, 523–527 (2008). both production and scavenging of reactive oxygen
hepatitis B virus and hepatitis C virus infection. Nature 42. Guo, J. T., Sohn, J. A., Zhu, Q. & Seeger, C. species. Oncology 81, S11–S17 (2011).
Rev. Immunol. 5, 215–229 (2005). Mechanism of the interferon α response against This report illustrates how HCV, through
18. Sun, B. S. et al. Hepatitis C virus replication in stably hepatitis C virus replicons. Virology 325, 71–81 impairment of mitochondrial function, results in
transfected HepG2 cells promotes hepatocellular (2004). oxidative stress, which is important to the
growth and tumorigenesis. J. Cell. Physiol. 201, 43. Li, K. et al. Immune evasion by hepatitis C virus pathogenesis of HCC.
447–458 (2004). NS3/4A protease-mediated cleavage of the Toll-like 64. Ripoli, M. et al. Hepatitis C virus-linked mitochondrial
19. Sells, M. A., Chen, M. L. & Acs, G. Production of receptor 3 adaptor protein TRIF. Proc. Natl Acad. Sci. dysfunction promotes hypoxia-inducible factor
hepatitis B virus particles in Hep G2 cells transfected USA 102, 2992–2997 (2005). 1α‑mediated glycolytic adaptation. J. Virol. 84,
with cloned hepatitis B virus DNA. Proc. Natl Acad. 44. Heim, M. H., Moradpour, D. & Blum, H. E. Expression 647–660 (2010).
Sci. USA 84, 1005–1009 (1987). of hepatitis C virus proteins inhibits signal 65. Quarato, G. et al. Targeting mitochondria in the
20. Fattovich, G., Stroffolini, T., Zagni, I. & Donato, F. transduction through the Jak-STAT pathway. J. Virol. infection strategy of the hepatitis C virus. Int.
Hepatocellular carcinoma in cirrhosis: incidence and 73, 8469–8475 (1999). J. Biochem. Cell Biol. 45, 156–166 (2012).
risk factors. Gastroenterology 127, S35–S50 (2004). 45. Polyak, S. J. et al. Hepatitis C virus nonstructural 5A 66. Merquiol, E. et al. HCV causes chronic endoplasmic
21. Guidotti, L. G. et al. Intracellular inactivation of the protein induces interleukin‑8, leading to partial reticulum stress leading to adaptation and
hepatitis B virus by cytotoxic T lymphocytes. Immunity inhibition of the interferon-induced antiviral response. interference with the unfolded protein response. PLoS
4, 25–36 (1996). J. Virol. 75, 6095–6106 (2001). ONE 6, e24660 (2011).
22. Akbar, S. M. F., Inaba. K. & Onji, M. Upregulation of 46. Taylor, D. R., Shi, S. T., Romano, P. R., Barber, G. N. & 67. Li, J. et al. Subversion of cellular autophagy machinery
MHC class II antigen on dendritic cells from hepatitis Lai, M. M. Inhibition of the interferon inducible protein by hepatitis B virus for viral envelopment. J. Virol. 85,
B virus transgenic mice by interferon‑γ abrogation of kinase PKR by HCV E2 protein. Science 285, 6319–6333 (2011).
immune response defect to a T‑cell-dependent antigen. 107–110 (1999). 68. Key, P. Y. & Chen, S. S. Activation of the unfolded
Immunology 87, 519–527 (1996). 47. Freeman, A. J. et al. The presence of an intrahepatic protein response and autophagy after hepatitis C virus
23. Yang, Y. F. et al. Interferon therapy in chronic hepatitis cytotoxic T lymphocyte response is associated with low infection suppresses innate antiviral immunity in vitro.
B reduces progression to cirrhosis and hepatocellular viral load in patients with chronic hepatitis C virus J. Clin. Invest. 121, 37–56 (2011).
carcinoma: a meta-analysis. J. Viral Hepat. 16, infection. J. Hepatol. 38, 349–356 (2003). 69. Dreux, M., Gastaminza, P., Wieland, S. F. &
265–271 (2009). 48. Thimme, R. et al. Determinants of viral clearance and Chisari, F. V. The autophagy machinery is required to
24. Chow, W. C. et al. Re‑treatment with interferon alfa of persistence during acute hepatitis C virus infection. initiate hepatitis C virus replication. Proc. Natl Acad.
patients with chronic hepatitis C. Hepatology 27, J. Exp. Med. 194, 1395–1306 (2001). Sci. USA 33, 14046–14051 (2009).
1144–1148 (1998). 49. Jahan, S., Ashfaq, U. A., Khaliq, S., Samreen, B. & 70. Ogata, M. et al. Autophagy is activated for cell survival
25. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: Afzal, N. Dual behavior of HCV core gene in regulation after endoplasmic reticular stress. Mol. Cell. Biol. 24,
the next generation. Cell 144, 646–674 (2011). of apoptosis is important in progression of HCC. Infect. 9220–9231 (2006).
26. Levrero, M. et al. Control of cccDNA function in Genet. Evol. 12, 236–239 (2012). 71. Liu, T. T. et al. A case-control study of the relationship
hepatitis B virus infection. J. Hepatol. 51, 581–592 50. Ng, S. A. & Lee, C. Hepatitis B virus X gene and between hepatitis B virus DNA level and risk of
(2009). hepatocarcinogenesis. J. Gastroenterol. 46, 974–990 hepatocellular carcinoma in Qidong, China. World
27. Jin, Y. M., Yun, C., Park, C., Wang, H. J. & (2011). J. Gastroenterol. 14, 3059–3063 (2008).
Cho, H. Expression of hepatitis B virus X protein is 51. Pan, J., Lian, Z., Wallett, S. & Feitelson, M. A. The 72. Vrancken, K., Poeshuyse, J. & Liekens, S. Angiogenic
closely correlated with the high periportal hepatitis B x antigen effector, URG7, blocks tumour activity of hepatitis B and C viruses. Antivir. Chem.
inflammatory activity of liver diseases. J. Viral Hepat. necrosis factor α-mediated apoptosis by activation of Chemother. 22, 159–170 (2012).
8, 322–330 (2001). phosphoinositol 3‑kinase and β-catenin. J. Gen. Virol. 73. Moon, E. J. et al. Hepatitis B virus X protein induces
28. Carreno, V., Bartolome, J., Castillo, I. & 88, 3275–3285, 2007. angiogenesis by stabilizing hypoxia-inducible
Quiroga, J. A. Occult hepatitis B virus and hepatitis C 52. Nakamura, H., Aoki, H., Hino, O. & Moriyama, M. HCV factor‑1α. FASEB J. 18, 382–384 (2004).
virus infections. Rev. Med. Virol. 18, 139–157 core protein promotes heparin binding EGF-like 74. Yoo, Y. G. et al. Hepatitis B virus X protein enhances
(2008). growth factor expression and activates Akt. Hepatol. transcriptional activity of hypoxia-inducible factor‑1α
29. Chen, S. Y. et al. Mechanisms for inhibition of hepatitis Res. 41, 455–462 (2011). through activation of mitogen activated protein kinase
B virus gene expression and replication by hepatitis C 53. Jiang, Y. F. et al. The oncogenic role of NS5A of pathway. J. Biol. Chem. 278, 39076–39084 (2003).
virus core protein. J. Biol. Chem. 278, 591–607 hepatitis C virus is mediated by up‑regulation of 75. Keith, B. & Simon M. C. Hypoxia-inducible facgtors,
(2003). survivin gene expression in the hepatocellular cell stem cells, and cancer. Cell 129, 465–472 (2007).
30. Lin, L., Verslype, C., van Pelt, J. F., van Ranst, M. & through p53 and NF‑κB pathways. Cell Biol. Int. 35, 76. Sanz-Camern, P. et al. Hepatitis B virus promotes
Fevery, J. Viral interaction and clinical implications of 1225–1232 (2011). angiopoietin‑2 expression in liver tissue. Role of HBV X
coinfection of hepatitis C virus with other hepatitis 54. Jin, Y. et al. The immune reactivity role of HCV induced protein. Am. J. Pathol. 169, 1215–1222 (2006).
viruses. Eur. J. Gastroenterol. Hepatol. 18, liver infiltrating lymphocytes in hepatocellular damage. 77. Abe, M. et al. Hepatitis C virus core protein
1311–1319 (2006). J. Clin. Immunol. 17, 140–153 (1997). up‑regulates the expression of vascular endothelial
31. Tamori, A. et al. Frequent detection of hepatitis B virus 55. Vandermeeren, A. M. et al. Subcellular forms and growth factor via the nuclear factor‑κB/hypoxia-
DNA in hepatocellular carcinoma of patients with biochemical events triggered in human cells by HCV inducible factor‑1α axis under hypoxic conditions.
sustained virologic response for hepatitis C virus. polyprotein expression from a viral vector. Virol. J. 5, Hepatol. Res. 42, 591–600 (2012).
J. Med. Virol. 81, 1009–1014 (2009). 102–121 (2008). 78. Schiffer, E. et al. Gefitinib, an EGFR inhibitor, prevents
32. Liu, Z. & Hou, J. Hepatitis B virus (HBV) and hepatitis 56. Clippinger, A. J. & Bouchard, M. J. Hepatitis B virus hepatocellular carcinoma development in the rat liver
C virus (HCV) dual infection. Int. J. Med. Sci. 3, 57–62 HBx protein localizes to mitochondria in primary rat with cirrhosis. Hepatology 41, 307–314 (2005).
(2006). hepatocytes and modulates mitochondrial membrane 79. Liu, J. et al. Increased expression of c‑erbB‑2 in liver is
33. Isogawa, M., Robek, M. D., Furuichi, Y. & potential. J. Virol. 82, 6798–6811 (2008). associated with HBxAg expression and shorter
Chisari, F. V. Toll-like receptor signaling inhibits 57. Bhargava, A. et al. Occult hepatitis C virus elicits survival in patients with hepatocellular carcinoma. Int.
hepatitis B virus replication in vivo. J. Virol. 79, mitochondrial oxidative stress in lymphocytes and J. Cancer 125, 1894–1901 (2009).
7269–7272 (2005). triggers PI3‑kinase-mediated DNA damage 80. El Bassuoni, M. A., Talaat, R. M., Ibrahim, A. A. &
34. Wu, J. et al. Hepatitis B virus suppresses toll-like response. Free Radic. Biol. Med. 51, 1806–1814 Shaker, O. T. TGF‑β1 and C‑erb‑B2 neu oncoprotein in
receptor-mediated innate immune responses in (2011). Egyptian HCV related chronic liver disease and
murine parenchymal and nonparenchymal liver cells. 58. Cho, H. K., Cheong, K. J., Kim, H. Y. & hepatocellular carcinoma patients. Egypt J. Immunol.
Hepatology 49, 1132–1140 (2009). Cheong, J. Endoplasmic reticulum stress induced by 15, 39–50 (2008).
35. Beg, A. A. & Baltimore, D. An essential role for NF‑κB hepatitis B virus X protein enhances cyclo-oxygenase 81. Collado, M. et al. Tumour biology: senescence in
in preventing TNF‑α‑induced cell death. Science 274, 2 expression via activating transcription factor 4. premalignant tumours. Nature 436, 642 (2005).
782–784 (1996). Biochem. J. 435, 431–439 (2011). 82. Benn, J. & Schneider, R. J. Hepatitis B virus HBx
36. Su, F. & Schneider, R. J. Hepatitis B virus HBx protein 59. Yang, F. et al. Expression of hepatitis B virus proteins protein activates Ras-GTP complex formation and
activates transcription factor NF‑κB by acting on in transgenic mice alters lipid metabolism and induces establishes a Ras, Raf, MAP kinase signaling cascade.
multiple cytoplasmic inhibitors of rel-related proteins. oxidative stress in the liver. J. Hepatol. 48, 12–19 Proc. Natl Acad. Sci. USA 91, 10350–10354 (1994).
J. Virol. 70, 4558–4566 (1996). (2008). 83. Oishi, N. et al. Hepatitis B virus X protein overcomes
37. Zhang, Z. et al. Hypercytolytic activity of hepatic 60. Srisuttee, R. et al. Up‑regulation of Foxo4 mediated by oncogenic ras-induced senescence in human
natural killer cells correlates with liver injury in chronic hepatitis B virus X protein confers resistance to immortalized cells. Cancer Sci. 98, 1540–1548 (2007).
hepatitis B patients. Hepatology 53, 73–85 (2011). oxidative stress-induced cell death. Int. J. Mol. Med. 84. Tanaka, S. & Shigeki, A. Molecular targeted therapy
38. van der Molen, R. G. et al. Functional impairment of 28, 255–260 (2011). for hepatocellular carcinoma in the current and
myeloid and plasmacytoid dendritic cells of patients 61. Yang, B. & Bouchard, M. J. The hepatitis B virus X potential next strategies. J. Gastroenterol. 46,
with chronic hepatitis B. Hepatology 40, 738–746 protein elevates cytosolic calcium signals by 289–296 (2011).
(2004). modulating mitochondrial calcium uptake. J. Virol. 86, This work discusses oncogene addiction in HCC as a
39. Yoo, Y. D. et al. Regulation of transforming growth 313–327 (2012). way to develop specific therapeutics aimed at
factor-β 1 expression by the hepatitis B virus (HBV) X 62. Bouchard, M. J., Puro, R. J., Wang, L. & pathways that are rate-limiting in tumorigenesis.
transactivator. Role in HBV pathogenesis. J. Clin. Schneider, R. J. Activation and inhibition of cellular 85. Chung, T. W., Lee, Y. C. & Kim, C. H. Hepatitis B viral
Invest. 97, 388–395 (1996). calcium and tyrosine kinase signaling pathways HBx induces MMP‑9 gene expression through
40. Lopes, A. R. et al. Bim-mediated deletion of antigen- identify targets of the HBx protein involved in activation of ERK and PI‑3K/AKT pathways:
specific CD8 T cells in patients unable to control HBV hepatitis B virus replication. J. Virol. 77, 7713–7719 involvement of invasive potential. FASEB J. 18,
infection. J. Clin. Invest. 118, 1835–1845 (2008). (2003). 1123–1125 (2004).

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 133

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

86. Tarn, C., Lee, S., Hu, Y., Ashendel, C. & 108. Kalluri, R. EMT: when epithelial cells decide to become 131. Monk, M. & Holding, C. Human embryonic genes
Andrisani, O. M. Hepatitis B virus X protein mesenchymal-like cells. J. Clin. Invest. 119, re‑expressed in cancer cells. Oncogene 20,
differentially activates RAS-RAF-MAPK & JNK 1417–1419 (2009). 8085–8091 (2001).
pathways in X‑transforming versus non-transforming 109. Castilla, A., Prieto, J. & Fausto, N. Transforming 132. Yamashita, T., Budhu, A., Forgues, M. &
AML12 hepatocytes. J. Biol. Chem. 276, growth factors β 1 and α in chronic liver disease. Wang, X. W. Activation of hepatic stem cell marker
34671–34680 (2001). Effects of interferon alfa therapy. N. Engl. J. Med. 324, EpCAM by Wnt‑β‑catenin signaling in hepatocellular
87. Hoare, M., Das, T. & Alexander, G. Ageing, telomeres, 933–940 (1991). carcinoma. Cancer Res. 67, 10831–10839 (2007).
senescence, and liver injury. J. Hepatol. 53, 950–961 110. Barrallo-Gimeno, A. & Nieto, M. A. The Snail genes as 133. Schoenhals, M. et al. Embryonic stem cell markers
(2010). inducers of cell movement survival: implications in expression in cancers. Biochem. Biophys. Res.
88. Kang, T. W. et al. Senescence surveillance of pre- development and cancer. Development 132, Commun. 383, 157–162 (2009).
malignant hepatocytes limits liver cancer 3151–3161 (2005). 134. Ali, N. et al. Hepatitis C virus-induced cancer stem cell-
development. Nature 479, 547–551 (2011). 111. van Ziji, F. et al. Epithelial-mesenchymal transition in like signatures in cell culture and murine tumor
89. Huang, J., Wang, Y., Guo, Y. & Sun, S. Down-regulated hepatocellular carcinoma. Future Oncol. 6, xenografts. J. Virol. 85, 12292–12303 (2011).
microRNA‑152 induces aberrant DNA methylation in 1169–1179 (2009). 135. Katoh, M. & Katoh, M. WNT signaling pathway and
hepatitis B virus-related hepatocellular carcinoma by 112. Lee, D. K. et al. The hepatitis B virus encoded stem cell signaling network. Clin. Cancer Res. 13,
targeting DNA methyltransferase 1. Hepatology 52, oncoprotein pX amplifies TGF‑β family signaling 4042–4045 (2007).
60–70 (2010). through direct interaction with Smad4: potential 136. Lu, T. Y. et al. Epithelial cell adhesion molecule
90. Benegiamo, G. et al. DNA methyltransferases 1 and mechanism of hepatitis B virus-induced liver fibrosis. regulation is associated with the maintenance of the
3b expression in Huh‑7 cells expressing HCV core Genes Dev. 15, 455–466 (2001). undifferentiated phenotype of human embryonic stem
protein of different genotypes. Dig. Dis. Sci. 57, 113. Xu, J. et al. TGF‑β‑induced epithelial-to‑mesenchymal cells. J. Biol. Chem. 285, 8719–8732 (2010).
1598–1603 (2012). transition. Cell Res. 19, 156–172 (2009). 137. Yamashita, T. et al. EpCAM-positive hepatocellular
91. Toyota, M. & Suzuki, H. Epigenetic drivers of genetic 114. Battaglia, S. et al. Liver cancer-derived hepatitis C carcinoma cells are tumor-initiating cells with stem/
alterations. Adv. Genet. 70, 309–323 (2010). virus core proteins shift TGF-β responses from tumor progenitor cell features. Gastroenterology 136,
92. Kao, C. F., Chen, S. Y., Chen, J. Y. & Wu suppression to epithelial-mesenchymal transition. 1012–1024 (2009).
Lee, Y. H. Modulation of p53 transcription regulatory PLoS ONE 4, e4355 (2009). This study shows that EpCAM functionally defines
activity and post-translational modification by 115. Zhu, Q. et al. miR‑21 promotes migration and invasion CSCs in HCC.
hepatitis C virus core protein. Oncogene 23, by the miR‑21‑PDCD4‑AP‑1 feedback loop in human 138. Ji, J. et al. Identification of microRNA‑181 by genome-
2472–2483 (2004). hepatocellular carcinoma. Oncol. Rep. 27, wide screening as a critical player in EpCAM-positive
93. Wang, X. W. et al. Hepatitis B virus X protein inhibits 1660–1668 (2012). hepatic cancer stem cells. Hepatology 50, 472–480
p53 sequence-specific DNA binding, transcriptional 116. Wu, C. Y., Tsai, Y. P., Wu, M. Z., Teng, S. C. & (2009).
activity, and association with transcription factor Wu, K. J. Epigenetic reprogramming and post- 139. Li, Z. & Rich, J. N. Hypoxia and hypoxia inducible
ERCC3. Proc. Natl Acad. Sci. USA 91, 2230–2234 transcriptional regulation during the epithelial- factors in cancer stem cell maintenance. Curr. Top.
(1994). mesenchymal transition. Trends Genet. 28, 454–463 Microbiol. Immunol. 345, 21–30 (2010).
94. Zhao, J. et al. Epigenetic silence of ankyrin-repeat- (2012). 140. Kaidi, A., Williams, A. C. & Paraskeva, G. Interaction
containing, SH3‑domain-containing, and proline-rich- This report highlights the importance of epigenetic between β-catenin and HIF‑1 promotes cellular
region-containing protein 1 (ASPP1) and ASPP2 genes events to EMT. adaptation to hypoxia. Nature Cell Biol. 9, 210–217
promotes tumor growth in hepatitis B virus-positive 117. Shih, W. L., Kuo, M. L., Chuang, S. E., Cheng, A. L. & (2007).
hepatocellular carcinoma. Hepatology 51, 142–153 Doong, S. L. Hepatitis B virus X protein activates a 141. Hoffmeyer, K. et al. Wnt/β‑catenin signaling regulates
(2010). survival signaling by linking SRC to telomerase in stem cells and cancer cells. Science
95. Machida, K. et al. Hepatitis C virus inhibits DNA phosphatidylinositol 3‑kinase. J. Biol. Chem. 278, 336, 1549–1554 (2012).
damage-repair through reactive oxygen and nitrogen 31807–31813 (2003). 142. Yang, J. & Weinberg, R. A. Epithelial-mesenchymal
species and by Interfering with the ATM‑NBS1/Mre11/ 118. Pfannkuche, A. et al. c‑Src is required for complex transition: at the crossroads of development and
Rad50 DNA repair pathway in monocytes and formation between the hepatitis C virus-encoded tumor metastasis. Dev. Cell. 14, 818–829 (2008).
hepatocytes. J. Immunol. 185, 6985–6998 (2010). proteins NS5A and NS5B: a prerequisite for 143. Copple, B. L. Hypoxia stimulates hepatocyte epithelial
96. Park, S. H., Jung, J. K., Lim, J. S., Tiwari, I. & replication. Hepatology 53, 1127–1136 (2011). to mesenchymal transition by hypoxia-inducible factor
Jang, K. L. Hepatitis B virus X protein overcomes all- 119. Yang, S. Z. et al. HBx protein induces EMT through and transforming growth factor-β-dependent
trans retinoic acid-induced cellular senescence by c‑Src activation in SMMC‑7721 hepatoma cell line. mechanisms. Liver Int. 30, 669–682 (2010).
down-regulating levels of p16 and p21 via DNA Biochem. Biophys. Res. Commun. 382, 555–560 144. Mani, S. A. et al. The epithelial-mesenchymal
methylation. J. Gen. Virol. 92, 1309–1317 (2011). (2009). transition generates cells with properties of stem cells.
97. Lim, J. S., Park, S. H. & Jang, K. L. Hepatitis C virus 120. Lara-Pezzi, E., Roche, S., Andrisani, O. M., Sánchez- Cell 133, 704–715 (2008).
core protein overcomes stress-induced premature Madrid, F. & López-Cabrera, M. The hepatitis B virus 145. Kurrey, N. K. et al. Snail and slug mediate
senescence by down-regulating p16 expression via HBx protein induces adherens junction disruption in a radioresistance and chemoresistance by antagonizing
DNA methylation. Cancer Lett. 321, 154–161 src-dependent manner. Oncogene 20, 3323–3331 p53‑mediated apoptosis and acquiring a stem-like
(2012). (2001). phenotype in ovarian cancer cells. Stem Cells 27,
98. Zender, L. et al. An oncogenomics-based in vivo RNAi 121. Arzumanyan, A. et al. Epigenetic repression of 2059–2068 (2009).
screen identifies tumor suppressors in liver cancer. Cell E‑cadherin expression by hepatitis B virus x antigen in 146. Yamashita, T. et al. EpCAM and a‑fetoprotein
135, 852–864 (2008). liver cancer. Oncogene 31, 563–572 (2012). expression defines novel prognostic subtypes of
99. Meek, D. W. Tumour suppression by p53: a role for 122. Akkari, L. et al. Hepatitis C viral protein NS5A induces hepatocellular carcinoma. Cancer Res. 68, 1451–1461
the DNA damage response? Nature Rev. Cancer 9, EMT and participates in oncogenic transformation of (2008).
714–723 (2009). primary hepatocyte precursors. J. Hepatol. 57, 147. Niu, D., Feng, H. & Chen, W. N. Proteomic analysis of
100. Gramantieri, L. et al. Cyclin G1 is a target of 1021–1028 (2012). HBV-associated HCC: Insights on mechanisms of
miR‑122a, a microRNA frequently down-regulated in 123. Ripoli, M. et al. Hypermethylated levels of E‑cadherin disease onset and biomarker discovery. J. Proteom.
human hepatocellular carcinoma. Cancer Res. 67, promoter in Huh‑7 cells expressing the HCV core 73, 1283–1290 (2010).
6092–6099 (2007). protein. Virus Res. 160, 74–81 (2011). 148. Feitelson, M. A. & Lee, J. Hepatitis B virus integration,
101. Hao, M., Zheng, S., Ding, H. & Huang, A. Regulation 124. Shin, J. Y. et al. HCV core protein promotes liver fragile sites, and hepatocarcinogenesis. Cancer Lett.
of microRNA‑122 on HBV replication by targeting HBx fibrogenesis via up‑regulation of CTGF with TGF‑β1. 252, 157–170 (2007).
sequence. Sheng Wu Yi Xue Gong Cheng Xue Za Zhi Exp. Mol. Med. 37, 138–145 (2005). 149. Hsu, S. et al. Essential metabolic, anti-inflammatory,
28, 784–789, 803 (2011) (in Chinese). 125. Ming‑Ju, H., Yih-Shou, H., Tzy-Yen, C. & Hui- and anti-tumorigenic functions of miR‑122 in liver.
102. Jangra, R. K., Yi, M. & Lemon, S. M. Regulation of Ling, C. Hepatitis C virus E2 protein induce reactive J. Clin. Invest. 122, 2871–2883 (2012).
hepatitis C virus translation and infectious virus oxygen species (ROS)-related fibrogenesis in the 150. Tsai, W. C. et al. MicroRNA‑122 plays a critical role in
production by the microRNA miR‑122. J. Virol. 84, HSC‑T6 hepatic stellate cell line. J. Cell. Biochem. 112, liver homeostasis and hepatocarcinogenesis. J. Clin.
6615–6625 (2010). 233–243 (2011). Invest. 122, 2884–2897 (2012).
103. Ura, S. et al. Differential microRNA expression 126. Oishi, N. & Wang, X. W. Novel therapeutic strategies 151. Bolondi, L. et al. Surveillance programme of cirrhotic
between hepatitis B and hepatitis C leading disease for targeting liver cancer stem cells. Int. J. Biol. Sci. 7, patients for early diagnosis and treatment of
progression to hepatocellular carcinoma. Hepatology 517–535 (2011). hepatocellular carcinoma: a cost effective analysis. Gut
49, 1098–10112 (2009). 127. Tong, C. M., Ma, S. & Guan, X. Y. Biology of hepatic 48, 251–259 (2001).
104. Guarino. M., Tosoni, A., Nebuloni, M. Direct cancer stem cells. J. Gastroenterol. Hepatol. 26, 152. Sterling, R. K. et al. Utility of Lens culinaris agglutinin-
contribution of epithelium to organ fibrosis: epithelial- 1229–1237 (2011). reactive fraction of α-fetoprotein and des-γ-carboxy
mesenchymal transition. Hum. Pathol. 40, 128. Tomuleasa, C. et al. Isolation and characterization of prothrombin, alone or in combination, as. biomarkers
1365–1376 (2009). hepatic cancer cells with stem-like properties from for hepatocellular carcinoma. Clin. Gastroenterol.
105. Bedossa, P. & Paradis, V. Liver extracellular matrix hepatocellular carcinoma. J. Gastrointestin. Liver Dis. Hepatol. 7, 104–113 (2009).
in health and disease. J. Pathol. 200, 504–515 19, 61–67 (2010). 153. Bertino, G. et al. Hepatocellualar carcinoma serum
(2003). 129. Arzumanyan, A. et al. Does the hepatitis B antigen markers. Semin. Oncol. 39, 410–433 (2012).
106. Gressner, O. A. Weiskirchen, R. & Gressner, A. HBx promote the appearance of liver cancer stem 154. Uemura, M. et al. Identification of the antigens
M. Evolving concepts of liver fibrogenesis provide new cells? Cancer Res. 71, 3701–3708 (2011). predominantly reacted with serum from patients with
diagnostic and therapeutic options. Comp. Hepatol. 6, This shows that HBx contributes to HCC through hepatocellular carcinoma. Cancer 97, 2474–2479
7 (2007). activation of stemness in the liver. (2003).
107. Zeisberg, M. et al. Fibroblasts derive from 130. Zhao, R. C., Zhu, Y. S. & Shi, Y. New hope for cancer 155. Anders, R. A. et al. cDNA microarray analysis of
hepatocytes in liver fibrosis via epithelial to treatment: exploring the distinction between normal macroregenerative and dysplastic nodules in end-
mesenchymal transition. J. Biol. Chem. 282, adult stem cells and cancer stem cells. Pharmacol. stage hepatitis C virus-induced cirrhosis. Am.
23337–23347 (2007). Ther. 119, 74–82 (2008). J. Pathol. 162, 991–1000 (2003).

134 | FEBRUARY 2013 | VOLUME 13 www.nature.com/reviews/cancer

© 2013 Macmillan Publishers Limited. All rights reserved


REVIEWS

156. Paradis, V. et al. Molecular profiling of hepatocellular 167. Martin, M. P. et al. IL28B polymorphism does not 177. Grant, T. J. et al. Antiproliferative small-molecule
carcinomas (HCC) using a large-scale real-time RT‑PCR determine outcomes of hepatitis B virus or HIV inhibitors of transcription factor LSF reveal oncogene
approach. Determination of a molecular diagnostic infection. J. Infect. Dis. 202, 1749–1753 (2010). addiction to LSF in hepatocellular carcinoma. Proc.
index. Am. J. Pathol. 163, 733–741 (2003). 168. Qi, P. et al. -509C>T polymorphism in the TGF‑β1 Natl Acad. Sci. USA 109, 4503–4508 (2012).
157. Lee, C. F., Ling, Z. Q., Zhao, T. & Lee, K. R. Distinct gene promoter, impact on the hepatocellular 178. Feitelson, M. A., Lian, Z., Liu, J., Tufan, N. L. &
expression patterns in hepatitis B virus- and hepatitis carcinoma risk in Chinese patients with chronic Pan, J. Parallel epigenetic and genetic changes in
C virus-infected hepatocellular carcinoma. hepatitis B virus infection. Cancer Immunol. hepatitis B virus associated hepatocellular carcinoma.
World J. Gastroenterol. 14, 6072–6077 (2008). Immunother. 58, 1433–1440 (2009). Cancer Lett. 239, 10–20 (2006).
158. Maass, T. et al. Microarray-based gene expression 169. Okamoto, K. et al. The genotypes of IL‑1 β and MMP‑3 179. Baylin, S. B. & Ohm, J. E. Epigenetic gene silencing
analysis of hepatocellular carcinoma. Curr. Genomics are associated with the prognosis of HCV-related in cancer – a mechanism for early oncogenic
11, 261–268 (2010). hepatocellular carcinoma. Intern. Med. 49, 887–895 pathway addiction? Nature Rev. Cancer 6, 107–116
159. Sun, M. et al. Expression profile reveals novel (2010). (2006).
prognostic biomarkers in hepatocellular carcinoma. 170. Llovet, J. M. & Bruix, J. Molecular targeted therapies This review provides a model that describes the
Front. Biosci. (Elite Ed) 2, 829–840 (2010). in hepatocellular carcinoma. Hepatology 48, relationship between epigenetic modifications in
160. Marquardt, J. U., Galle, P. R. & Teufel, A. Molecular 1312–1327 (2008). gene expression and mutations in these same
diagnosis and therapy of hepatocellular carcinoma This is an early report outlining the importance of genes in cancer pathogenesis.
(HCC): an emerging field for advanced technologies. targeting pathways that are operative in the 180. Popovic, R. & Licht, J. D. Emerging epigenetic targets
J. Hepatol. 56, 267–275 (2012). pathogenesis of HCC. and therapies in cancer medicine. Cancer Discov. 2,
161. Abu Dayyeh, B. K. et al. A functional polymorphism in 171. Menzo, S. et al. Trans-activation of epidermal growth 405–413 (2012).
the epidermal growth factor gene is associated with factor receptor gene by the hepatitis B virus X‑gene 181. Tiollais, P., Pourcel, C. & Dejean, A. The hepatitis B
risk for hepatocellular carcinoma. Gastroenterology product. Virology 196, 878–882 (1993). virus. Nature 317, 489–495 (1985).
141, 141–149 (2011). 172. Diao, J. et al. Hepatitis C virus induces epidermal 182. Rehermann, B. Hepatitis C virus versus innate and
162. Tanabe, K. E. Epidermal growth factor gene functional growth factor receptor activation via CD81 binding for adaptive immune responses: a tale of coevolution
polymorphism and the risk of hepatocellular viral internalization and entry. J. Virol. 86, and coexistence. J. Clin. Invest. 119, 1745–1754
carcinoma in patients with cirrhosis. JAMA 299, 10935–10949 (2012). (2009).
53–60 (2008). 173. Sedlaczek, N., Hasilik, A., Neuhaus, P., Schuppan, D. &
163. Thomas, D. L. et al. Genetic variation in IL28B and Herbst, H. Focal overexpression of insulin-like growth Acknowledgements
spontaneous clearance of hepatitis C virus. Nature factor 2 by hepatocytes and cholangiocytes in viral liver This work was supported by US National Institutes of Health
461, 798–801 (2009). cirrhosis. Br. J. Cancer 88, 733–739 (2003). (NIH) grants AI076535 and CA104025 to M.A.F.
164. Ge, D. et al. Genetic variation in IL28B predicts 174. Liu, J. et al. Enhancement of canonical Wnt/β‑catenin
hepatitis C treatment-induced viral clearance. Nature signaling activity by HCV core protein promotes cell Competing interests statement
461, 399–401 (2009). growth of hepatocellular carcinoma cells. PLoS ONE 6, The authors declare no competing financial interests.
165. Qi, P. et al. No association of EGF 5’UTR variant A61G e27496 (2011).
and hepatocellular carcinoma in Chinese patients with 175. Blivet-Van Eggelpoël, M. J. et al. Epidermal growth
chronic hepatitis B virus infection. Pathology 41, factor receptor and HER‑3 restrict cell response to FURTHER INFORMATION
555–560 (2009). sorafenib in hepatocellular carcinoma cells. J. Hepatol. Mark A. Feitelson’s homepage: http://www.temple.edu/cst/
166. Li, W. et al. Expression and gene polymorphisms of 57, 108–115 (2012). research/profiles/biology/feitelson.html
interleukin 28B and hepatitis B virus infection in a 176. Villanueva, A. & Llovet, J. M. Targeted therapies for ClinicalTrials.gov: http://clinicaltrials.gov/
Chinese Han population. Liver Int. 31, 1118–1126 hepatocellular carcinoma. Gastroenterology 140, ALL LINKS ARE ACTIVE IN THE ONLINE PDF
(2011). 1410–1426 (2011).

NATURE REVIEWS | CANCER VOLUME 13 | FEBRUARY 2013 | 135

© 2013 Macmillan Publishers Limited. All rights reserved

You might also like